Non-unitary Wightman CFTs and non-unitary vertex algebras
Abstract
We give an equivalence of categories between: (i) Möbius vertex algebras which are equipped with a choice of generating family of quasiprimary vectors, and (ii) (not-necessarily-unitary) Möbius-covariant Wightman conformal field theories on the unit circle. We do not impose any technical restrictions on the theories considered (such as finite-dimensional conformal weight spaces or simplicity), yielding the most general equivalence between these two axiomatizations of two-dimensional chiral conformal field theory. This provides new opportunities to study non-unitary vertex algebras using the lens of algebraic conformal field theory and operator algebras, which we demonstrate by establishing a non-unitary version of the Reeh-Schlieder theorem.
1 Introduction
It is a fundamental mathematical challenge to establish a rigorous axiomatization of quantum field theory (QFT), and in general this problem remains wide open except in very specialized contexts. In recent years, axiomatic QFT has received particular attention in the context of two-dimensional chiral conformal field theories (CFTs), as these theories are sufficiently structured to enable a rigorous mathematical treatment while at the same time exhibiting a wide variety of mathematical connections (to operator algebras and subfactors, to representation theory and modular tensor categories, to vector-valued modular forms, and to many other areas). There are many proposed axiomatizations of two-dimensional chiral CFTs, each of which captures different aspects of the physical theory, and none of which have been rigorously demonstrated to provide a complete description of CFT. It is conjectured that these different axiomatizations are essentially equivalent, and there have been recent breakthroughs in comparing different axiomatizations under certain technical hypotheses [CKLW18, Ten19a].
In this article we demonstrate the equivalence of two well-known axiomatizations of two-dimensional chiral CFTs. We establish this equivalence in the most general way possible, without any reliance on auxiliary technical hypotheses or restrictions on the models under consideration, such as the existence of an invariant inner product (“unitarity”). Proving equivalences at this level of generality has largely been viewed as an aspirational (but not necessarily feasible) goal of axiomatic QFT, which we achieve here through a detailed analysis of the mathematical structures in question.
The first axiomatization that we consider is the non-unitary version of the (bosonic) Wightman axioms on the unit circle , with Möbius symmetry (i.e. symmetry group , the holomorphic automorphisms of the unit disk). The key data of such a theory is a collection of operator-valued distributions (or Wightman fields) acting on a common invariant vector space of states , along with a compatible positive-energy representation of .
The second axiomatization that we consider is (-graded, bosonic) Möbius vertex algebras. These are vertex algebras graded by non-negative integer conformal dimensions, with symmetry given by the complexified Lie algebra .
We prove the following main result.
Main Result.
There is a natural equivalence of categories between non-unitary Möbius-covariant Wightman conformal field theories on and Möbius vertex algebras equipped with a family of quasiprimary generators.
Our result does not require unitarity or the existence of an invariant bilinear form, and we do not require that the homogeneous subspaces for the grading by conformal dimensions be finite-dimensional. There are many important examples of CFTs arising in mathematical and theoretical physics which require this level of generality, and in particular non-unitarity arises from the CFT-approach to classical critical phenomena, and from string theory. Specific examples include the non-unitary Virasoro minimal models, affine vertex algebras at non-critical level (both universal and simple quotient), bosonic -graded affine W-algebras (again universal and simple quotient), and the -ghost vertex algebra with central charge (along with other “-graded” vertex algebras which arise in logarithmic conformal field theory). As a result of our theorem there are canonical Wightman CFTs associated to these models, which demonstrates significant functional analytic regularity that is not otherwise apparent.
The constructions going from Wightman CFTs to vertex algebras and back are given in Sections 3.1 and 3.2, respectively. These are shown to give an equivalence of categories in Section 3.3. The vertex algebra associated to a Wightman CFT with domain is constructed as a certain subspace . Conversely, the Wightman CFT associated to is constructed as an extension . We note that at this level of generality (allowing each weight space to be infinite-dimensional) it is not a priori clear that there is a single Wightman CFT for each vertex algebra, and it seems plausible that there could be families of ‘Wightman completions’ of a single vertex algebra. However, as a consequence of our result, there is indeed a unique Wightman CFT for each Möbius vertex algebra.
A very useful and inspiring heuristic discussion on the connection between Wightman CFTs and vertex algebras can be found in [Kac98, §1.2]. However, the arguments given there do not appear to be aimed to give precise mathematical details on this connection. More recently, three of the present authors gave a rigorous proof that unitary Möbius vertex algebras were equivalent to unitary Wightman CFTs possessing an additional analytic property called uniformly bounded order, provided that the homogeneous subspaces for the grading by conformal dimensions were finite-dimensional [RTT22]. The present article generalizes the previous result to possibly non-unitary theories, also dropping the requirements of uniformly bounded order and finite-dimensional weight spaces. As the techniques historically used to study Wightman theories involve careful analysis of the norm topology on the space of states, there is significant new work required to generalize our previous results to the non-unitary setting.
We also demonstrate in Section 4 that the correspondence constructed in this article is compatible with invariant bilinear forms, invariant sesquilinear forms, and invariant inner products.
Theorem.
Let be a Möbius-covariant Wightman CFT and let be the associated Möbius vertex algebra. Then every invariant inner product (unitary structure) on restricts to an invariant inner product on , and conversely every unitary structure on uniquely extends to one on . The same holds for invariant sesquilinear forms (involutive structures) and invariant bilinear forms.
We are left with a striking and clear correspondence between two well-known axiomatizations of two-dimensional chiral conformal field theory, without any reliance on additional technical hypotheses. We are motivated in part by the possibility to provide such an equivalence, which is not often possible in the wild landscape of axiomatic quantum field theory. We are also motivated by intriguing links between non-unitary conformal field theories and the unitary world of algebraic conformal field theory. Given a Wightman CFT on and an interval , consider the algebra generated by Wightman fields smeared by test functions supported in the interval . Such Wightman nets of algebras have been studied in the context of unitary quantum field theories [SW64], and there is a substantial effort underway to understand the relationship between unitary vertex algebras, unitary Wightman nets, and the usual nets of algebras of bounded observables (i.e. conformal nets) studied in algebraic conformal field theory [CKLW18, Ten19a]. On the other hand, as a result of our present work, there exists a Wightman net for every Möbius vertex algebra, including non-unitary ones. Such nets could give an avenue to apply methods generally used in the unitary framework of algebraic quantum field theory in the more general setting of non-unitary models. Previously such links have been probed only at the level of categories of representations [EG17]. As a first demonstration of the potential of this approach we prove a version of the Reeh-Schlieder Theorem (regarding the cyclic and separating property of the vacuum vector) for non-unitary theories in Appendix A.
Finally, there is strong motivation to understand functional analytic aspects of non-unitary vertex algebras as a part of studying links between algebraic and geometric aspects of the theory, as in [Hua99, Hua03]. More recently, analytic considerations of non-unitary vertex algebras have played a key role in the study of conformal blocks [Gui24a, Gui24b, GZ23], and such considerations also feature in the construction of functorial CFTs in the sense of Segal [Seg04].
In future work, it would be interesting to relate modules for vertex algebras to representations of the corresponding Wightman nets, which would fit into the broad program underway in the unitary setting to relate vertex algebra modules to representations in algebraic conformal field theory [Ten19b, Ten24, Gui21, Gui20, CWX]. Such relations should enable further correspondences between full two-dimensional conformal field theories in various approaches, cf. [Mor23, AGT23, AMT24].
Acknowledgements
S.C. and Y.T acknowledge support from the GNAMPA-INDAM project Operator algebras and infinite quantum systems, CUP E53C23001670001 and from the MIUR Excellence Department Project MatMod@TOV awarded to the Department of Mathematics, University of Rome “Tor Vergata”, CUP E83C23000330006. C.R. and J.T. were supported by ARC Discovery Project DP200100067, “Physical realisation of enriched quantum symmetries”.
2 Preliminaries on Wightman CFTs and Möbius vertex algebras
An operator-valued distribution on the unit circle with domain a vector space is a linear map
where is the space of linear operators on . In this article, we will typically study operator-valued distributions whose domain is infinite-dimensional, and we will require some topological considerations with respect to the action of sets of such distributions on .
If is a set of operator-valued distributions on with a common domain , then a linear functional is called compatible with if the multilinear maps given by
(2.1) |
are continuous in the for all and . Note that multilinear forms are separately continuous if and only if they are jointly continuous since is a Fréchet space [Trè67, Cor. §34.2]. We write for the space of all linear functionals compatible with . Recall that a set of linear functionals on is said to separate points if for every non-zero there is a such that .
Definition 2.1.
A set of operator-valued distributions with domain acts regularly if separates points.
If we imagine that consists of a family of Wightman fields (i.e. the operators are smeared quantum fields), then it is natural that expectations should have the property that expressions (2.1) are continuous in the smearing functions . Thus, the condition of regularity serves to exclude certain nonphysical actions that have the property that expectations cannot distinguish states. The following example illustrates the pathological behavior of nonregular actions.
Example 2.2.
Let be the tensor algebra, and for let be the operation of left-multiplication by in . The space carries a regular action of . Let be the left ideal generated by trigonometric polynomials . Let , and observe that for each the action of descends to an operator . The action of on is not regular. Let be the unit of the tensor algebra, and let be its image under the canonical projection. For any we have , and thus any vanishes on for any . In particular, if , then is non-zero but lies in the kernel of all .
Remark 2.3.
A non-regular action of on descends to a regular action on the quotient , where .
Let be the function . The condition that acts regularly on ensures that the operators are determined by the modes in a certain sense that we will make precise below. This is in contrast with Example 2.2, in which for all but for some .
We now introduce certain topologies on associated with the action of . We assume here that the reader is familiar with (or indifferent to) the fundamentals of topological and locally convex vector spaces, and defer the relevant background and additional details to Appendix B.
Definition 2.4.
Given a family of operator-valued distributions on with domain , the -weak topology on is the weak topology induced by the linear functionals . That is, the -weak topology is the coarsest topology such that every is continuous.
For a topological vector space , a map is -weakly continuous precisely when is continuous for all . The family acts regularly precisely when the -weak topology is Hausdorff. We will see in Lemma 2.8 below that for the expressions are continuous in when is given the -weak topology, so indeed is determined by the modes when acts regularly.
There is a second natural topology on associated with the action of . For and , let
be the linear map
We equip the algebraic tensor product with the projective topology, for which continuous linear maps correspond to continuous multilinear maps (see Appendix B).
Definition 2.5.
Given a family of of operator-valued distributions on with domain , the -strong topology on is the colimit (or final) locally convex topology induced by the maps for and . That is, the -strong topology is the finest locally convex topology such that the maps are continuous.
Equivalently, the -strong topology is the finest locally convex topology on such that expressions are continuous in the functions (jointly, or equivalently separately by [Trè67, Cor. §34.2]).
Remark 2.6.
For a locally convex space , a linear map is continuous precisely when is continuous for all and [NB11, Thm. 12.2.2]. In particular, a linear functional is -strongly continuous if and only if , and so the weak topology on induced by the space of -strong continuous linear functionals is precisely the -weak topology.
We now have the following alternate characterizations of the regularity of an action of on .
Lemma 2.7.
Let be a set of operator-valued distributions on with domain a vector space . Then the following are equivalent.
-
i)
acts regularly on , i.e. separates points.
-
ii)
The -weak topology on is Hausdorff.
-
iii)
The -strong topology on is Hausdorff.
-
iv)
There exists a locally convex Hausdorff topology on such that the maps
are continuous for all and .
Proof.
As noted above, the implication (i) (ii) follows immediately from the definitions of regularity and the -weak topology. The identity map is continuous from the -weak topology to the -strong topology, and thus (ii) (iii). The -strong topology is locally convex by definition, and thus (iii) (iv) is tautological. Finally, if is a locally convex Hausdorff topology on as in (iv), then we have an inclusion of continuous duals . By the Hahn-Banach theorem separates points [NB11, Thm. 7.7.7]. Hence so does , and the action of is regular. ∎
Both the -strong and -weak topologies are quite natural, and so it is not surprising that the fields act continuously when is given one of these topologies.
Lemma 2.8.
Let be a set of operator-valued distributions on acting regularly with domain equipped with the -strong topology. Then for the natural map is separately continuous. The same holds if is equipped with the -weak topology.
Proof.
First, fix . For any and , the expression
is continuous in the by the definition of the -strong topology, and thus is continuous by Remark 2.6. Similarly, for fixed expressions are continuous in , and we conclude that is separately continuous.
Now consider if is given the -weak topology. For the expression is evidently continuous in , and it just remains to show that acts continuously on . Let be the algebraic dual of , and let be the adjoint action. If then
(2.2) |
is continuous in the functions , so leaves invariant. Thus if is a net in converging -weakly to then
Hence converges -weakly to and acts continuously on . ∎
We now assume our vector space is equipped with a family of operator-valued distributions on as well as a representation , where is the group of holomorphic automorphisms of the closed unit disk. As in [CKLW18, §6], for we denote by the function
which takes positive real values since is an orientation-preserving diffeomorphism of . For and we denote by the function
(2.3) |
An operator-valued distribution with domain is called Möbius-covariant with conformal dimension under the representation if for every and every we have
as endomorphisms of . We say that a vector has conformal dimension if for all rotations .
We now present the not-necessarily-unitary version of the Wightman axioms for two-dimensional chiral conformal field theories on the circle . Historically, the Wightman axioms have been closely entwined with unitary theories, where the space of states possess an appropriate inner product. Non-unitary versions of the Wightman axioms have also appeared in various contexts such as the mathematical description of gauge fields, see e.g. [Str93, §6.4]. In this article we will generally refer to the non-unitary theories in question simply as Wightman conformal field theories for the sake of brevity.
Definition 2.9.
Let be a vector space equipped with a representation of and a choice of non-zero vector . Let be a set of operator-valued distributions on acting regularly on their common domain . This data forms a (not-necessarily-unitary) Möbius-covariant Wightman CFT on if they satisfy the following axioms:
-
(W1)
Möbius covariance: For each there is such that is Möbius-covariant with conformal dimension under the representation .
-
(W2)
Locality: If and have disjoint supports, then and commute for any pair .
-
(W3)
Spectrum condition: If has conformal dimension then .
-
(W4)
Vacuum: The vector is invariant under , and is spanned by vectors of the form .
A Möbius-covariant Wightman CFT is a quadruple , but we will frequently refer to the family of fields or the domain as a (Möbius-covariant) Wightman CFT when the remaining data is clear from context.
Let , and let
By Möbius covariance (W1) the vectors in have conformal dimension , or in other words acts on as multiplication by .
Lemma 2.10.
Let (,,,) be a Möbius-covariant Wightman CFT, and suppose is equipped with the -strong topology.
-
i)
The map is separately continuous.
-
ii)
is dense in .
The same holds for the -weak topology.
Proof.
First consider the -strong topology. Fix , and we will show that is continuous in . By the vacuum axiom we may assume without loss of generality that . We have
The smooth function depends continuously on , and thus depends continuously on as well. Now consider a fixed , and by the same argument depends continuously on the . Hence by Remark 2.6 and the vacuum axiom depends continuously on , proving (i). The argument is similar for the -weak topology.
For (ii), note that is dense since Laurent polynomials are dense in , and for by the spectrum condition. ∎
Let denote the space of linear functionals such that when . The following technical observations will be essential in constructing vertex algebras from Wightman CFTs.
Lemma 2.11.
Suppose that (,,,) satisfies all of the axioms of a Möbius-covariant Wightman CFT except perhaps for the spectrum condition.
-
i)
For , separates points in .
-
ii)
If is -strongly (or -weakly) dense then the spectrum condition holds.
Proof.
First we prove (i). From the definition of a Wightman CFT, separates points so given we may choose such that . For , let be rotation by . If is the adjoint operator, then . Let be given by
To see that the integral exists, observe that
Since is jointly continuous , and , the expression is jointly continuous , and thus the integral defining exists. Moreover, we see that . If , then , so and .
For part (ii), it suffices to consider the -strong topology. Let be the subspace of vectors with conformal dimension . Note that
but equality is not immediate (it is e.g. a consequence of Proposition 3.14). Now fix and let . To verify the spectrum condition we must show that . Let , and as above let . Then , but vanishes on . Since the latter space is assumed to be dense, and is continuous, we have . Hence , and since separates points we have , as desired. ∎
We will later see that for every there is a unique extension to a linear functional in that vanishes on (when ) – see Proposition 3.14.
The notion of Wightman CFT presented in Definition 2.9 generalizes the unitary notion of Wightman CFT considered in [RTT22] in several ways. Most notably, the domain does not have an inner product. We also do not assume that is the unique -invariant vector up to scale, and we do not require that the eigenspaces for the generator of rotation are finite-dimensional.
The requirement that act regularly on is necessary, as there exist pathological examples of quadruples that satisfy all of the requirements to be a Wightman CFT except for the regularity of the action of . Indeed, we can refine Example 2.2 to produce a quadruple with non-regular action for which the only finite-energy vectors are scalar multiples of the vacuum, despite being infinite-dimensional.
Example 2.12.
Let be the closed span of . These are the functions in that extend to holomorphic functions outside the unit disk which vanish at infinity. We similarly define . Let be the projection with kernel . Let be the symmetric algebra, and let be the unit. Let be the operator-value distribution with domain where acts by multiplication by in . This action is evidently regular. The family acts locally on since the symmetric algebra is abelian, and the vacuum vector is cyclic for .
Define a representation of on by , and extend this representation to (this representation is better understood as the quotient of the natural representation of on by constant functions). The representation has positive energy, and is -invariant. Moreover we have for
and it follows that . Hence is Möbius covariant with conformal dimension 1, and we have shown that is a Wightman CFT.
Now let be the dense subspace of functions which extend holomorphically to a neighborhood of . Then is invariant under . Moreover, a function lies in precisely when its Fourier coefficients decay sufficiently rapidly, and thus is invariant under as well. Let be the left ideal generated by , and observe that is invariant under and for all and . Let , let be the image of under the canonical projection, and let and be the operators on induced by and , respectively. The quadruple satisfy all of the requirements of a Wightman CFT except for regularity of the action of . We have
for all non-zero since .
Remark 2.13.
If satisfy all of the requirements of a Wightman CFT except that the fields do not act regularly, then we can obtain a Wightman CFT on a quotient of . Let . Since is invariant under the adjoint actions and , it follows that is invariant under and , and so we have actions of smeared fields and on the quotient . So long as , these actions give a Wightman CFT on .
If one applies this procedure to the example constructed in Example 2.12, then one obtains the trivial Wightman CFT. Indeed, we can check that as follows. Returning to the notation of Example 2.12, we have , and thus
Hence it suffices to check that an arbitrary vanishes on each of the spaces when . Note that this space is spanned by vectors of the form . When one of the functions lies in we have and thus . Hence for any we have
(2.4) |
when some lies in . Since is dense in and , it follows that (2.4) holds for arbitrary functions , and we conclude that vanishes on when . Hence as claimed, and is spanned by the vacuum .
With these distinctions in mind, we give the corresponding notion of vertex algebra after some brief preliminaries. Let be the three-dimensional real Lie algebra of . If is regarded as a subgroup of the group of orientation-preserving diffeomorphisms of the unit circle , then is identified with a three-dimensional subspace of the space of smooth vector fields on . Each vector field is identified with a differential operator for some smooth function , and the Lie bracket is given by , where denotes . Note that this bracket is the opposite of the bracket of vector fields, which is the natural choice when identifying with the Lie algebra of . The complexification of is spanned by the elements , where is the complexified vector field . The vector fields satisfy the commutation relations
In a representation of , we will frequently abuse notation and write for the operator corresponding to the vector field indicated above.
If is a vector space we write for the vector space of formal power series in with coefficients in . Given and , we have a formal series with coefficients in . For any we have . If is another formal series in a second formal variable , then the expression makes sense as a formal series in and .
We can now precisely specify the flavor of vertex algebras that we will consider.111The term Möbius vertex algebra has been used in the literature to describe various slightly different notions (see e.g. [BK08, HLZ14, Hua20, Kac98]). In some cases, authors include the possibility of fermionic fields, with the corresponding super version of the locality axiom; the term Möbius vertex superalgebra is also used in this case. Additionally, some authors replace our -grading with a more general grading. We do not foresee any significant obstacles to generalizing our results to Möbius vertex superalgebras graded by a lower-bounded subset of .
Definition 2.14.
An (-graded) Möbius vertex algebra consists of a vector space equipped with a representation of , a state-field correspondence , and a choice of non-zero vector such that the following hold:
-
(VA1)
, where .
-
(VA2)
and , i.e. has only non-negative powers of for all .
-
(VA3)
is -invariant, i.e. for .
-
(VA4)
and for all and .
-
(VA5)
for sufficiently large.
For , we write , where are called the modes of . A vector is called homogeneous (with conformal dimension ) if it lies in . As a consequence of the -commutation relation, when is homogeneous with conformal dimension we have . Hence maps into . A vector is called quasiprimary if it is homogeneous and .
We record two useful identities satisfied by the modes of a vertex operator (see [Kac98, §4.8]), the Borcherds product formula:
(2.5) |
and the Borcherds commutator formula:
(2.6) |
Note that when the sums of operators on the right-hand sides of (2.5) and (2.6) are applied to a vector, all but finitely many terms vanish.
3 Equivalence between Möbius vertex algebras and Wightman CFTs
3.1 From vertex algebras to Wightman CFTs
In this section we construct a Wightman CFT from a Möbius vertex algebra . The first step is to construct operator-valued distributions from the formal distributions , as follows. For , the degree-shifted mode is defined by , which gives an alternative field expansion , so that . We extend the definition of to non-homogeneous vectors by linearity. Let us write for the restricted dual , which is to say linear functionals on that are supported on finitely many . We denote by the algebraic completion
and we embed in the natural way. We equip with the weak topology induced by the pairing with .
For we define
by
where is the -th Fourier coefficient of .
We now show that the maps may be extended to act on an invariant domain . The first step is the following lemma, which has an identical proof to [RTT22, Lem. 2.7].
Lemma 3.1.
For all and , there exists a polynomial such that
for all . The polynomial depends on the vectors and , but the degree of may be bounded independent of and .
We refer the reader to [RTT22] for a proof. A very similar argument, however, yields the following observation.
Lemma 3.2.
Let and let . Then
is finite-dimensional.
Proof.
We proceed by induction on , with the cases and being immediate. Now fix , and suppose when . First observe that when and we must have , and hence
Next observe that
and each of the subspaces is finite-dimensional by the inductive hypothesis, so it suffices to show that the final term is finite-dimensional as well.
By the Borcherds commutator formula (2.6), if and we have
The subspace on the right-hand side is finite-dimensional by the inductive hypothesis and independent of , and thus we conclude that is finite-dimensional as well. ∎
If , then maps into . Our next lemma gives an estimate for these maps in terms of the -Sobolev norm of . Recall that for , the -Sobolev norm on is given by
(3.1) |
We denote by the Hilbert space completion of under this norm, which consists of -functions with finite -Sobolev norm. The locally convex topology on is induced by the norms , and a linear map from to a Banach space is continuous precisely when it is bounded with respect to some -Sobolev norm. We then have the following estimate, which is a simplification of [RTT22, Lem. 2.8].
Lemma 3.3.
For all , , and Laurent polynomials , we have
The number depends only on the , and the constant depends on the , , and .
The auxiliary domain and the topology on it.
By Lemma 3.2, the assignment gives a map
with each space finite-dimensional. By Lemma 3.3, this extends to a continuous multilinear map again taking values in . Thus for each , there exists a unique continuous multilinear map
such that when we have
(3.2) |
Let , and for set
We have by considering . Let , equipped with the subspace topology (i.e. the weak topology induced by the linear functionals , in which a sequence (or net) converges to if and only if converges to for all ).
Lemma 3.4.
For all and there exists a unique continuous linear map such that:
-
i)
.
-
ii)
The expressions are (jointly) continuous in the functions .
In addition, we have .
Proof.
We first consider uniqueness. When are Laurent polynomials, the condition determines the value of . The value of such expressions in is then uniquely determined by continuity in the functions .
We now show existence. We wish to define on by the formula
(3.3) |
but must check that this is well-defined.
First, consider if is a Laurent polynomial. The modes map to , and thus the adjoint (transpose) operator maps into . Hence there is an adjoint map such that for and we have
Thus if are Laurent polynomials we have
As the first and last terms are jointly continuous in by (3.2), we have
(3.4) |
whenever is a Laurent polynomial.
We now argue that (3.3) is well-defined. Let be the linear map corresponding to the multilinear map , so that is the range of . Let , and let be the map given by on the th direct summand of . We wish to show that if and , then as well.
Fix as above. By (3.4), if is a Laurent polynomial we have for all
and so . On the other hand, is continuous in , and so vanishes for all . Thus there is a well-defined map satisfying (3.3).
By construction we have . It follows immediately that such expressions span , and since is continuous in the functions we have also shown the second required property of the operator . For the first property, note that agrees with by (3.2) when is a Laurent polynomial, and thus for all by continuity. ∎
Remark 3.5.
We have constructed a family of operator-valued distributions on . We next consider Möbius covariance of these distributions, which will hold when is quasiprimary. To this end we introduce
Note that , so that every vector in may be written uniquely as a sum of homogeneous quasiprimary components. Let
Let us assume that is generated by as a vertex algebra, in which case . If is completely reducible as a -module (i.e. if it is spanned by vectors of the form with ), then it is evidently generated by as a vertex algebra and moreover as when .
We will now construct a representation of for which the Wightman fields are covariant for all and for which the vacuum is invariant. Note that such a representation is unique if it exists, as the covariance condition implies
So the difficulty is in showing that a linear map satisfying the above condition exists, i.e. showing that if a linear combination of vectors of the form
vanishes, then so does the corresponding linear combination of
We first extend the representation of furnished by the Möbius vertex algebra structure on to a representation on . Recall that is spanned by complexified vector fields on the circle, where corresponds to .
Let be the algebraic dual of , and note that the adjoint operators leave invariant. We claim that the closure of the graph in is the graph of a densely defined linear operator. Indeed, suppose that is a net in such that and in . Then for any it holds that
As separates points in , we conclude that and that the closure of is the graph of a densely defined operator as claimed. Taking linear combinations we obtain a densely-defined operator on for every , which we denote by .
Lemma 3.6.
Let be a Möbius vertex algebra that is generated as a vertex algebra by its quasiprimary fields. Then for any the domain of contains and leaves invariant. Moreover, if is quasiprimary with conformal dimension we have
as endomorphisms of .
Proof.
If then the commutation relations between and the from the definition of a Möbius vertex algebra imply that when is a Laurent polynomial we have
as endomorphisms of . Thus if are Laurent polynomials and , then we have
(3.5) |
where is the conformal dimension of . For arbitrary , choose sequences of Laurent polynomials222 and it is not the -th Fourier coefficient of . such that in , and observe that
and
in by Lemma 3.4. Hence lies in the domain of and (3.5) holds for . It follows that leaves invariant and we have the desired commutation relation with smeared fields. ∎
We now turn to constructing the desired representation . The following lemma will allow us to define on a .
Lemma 3.7.
Let be a Möbius vertex algebra which is generated by a set of quasiprimary vectors. Then for any there exists a unique linear map such that
for all with conformal dimensions and all .
Proof.
Uniqueness is clear as the required formula for determines its value on . In order to show existence of , we must show that if a linear combination of vectors of the form vanishes, then so does the corresponding linear combination of vectors of the form . We use standard ODE techniques.
The exponential map is surjective, so we may choose such that . Let be the corresponding one-parameter subgroup of . Let and consider the function given by
We will now show that extends holomorphically to a neighborhood of (when is given the weak topology induced by ), and compute its derivative.
The map given by extends holomorphically to a neighborhood of (taking values in complex Möbius transformations of the Riemann sphere ). For each , the Möbius transformation leaves invariant, and thus for a sufficiently small neighborhood of the corresponding Möbius transformations map into . Thus if is a Laurent polynomial, the function given by extends holomorphically to a neighborhood of . It follows that the map sending extends holomorphically to a neighborhood of .
Fix . By Lemma 3.4, the expressions are jointly continuous in . Thus for fixed Laurent polynomials , the function
extends holomorphically to a neighborhood of . As this neighborhood is independent of , the function
extends holomorphically to a neighborhood of , as previously claimed.
We now differentiate the above function of . A straightforward computation [RTT22, Eqn. (3.4)] shows that
(3.6) |
with the derivative taken in . Comparing (3.6) with the commutation relation of Lemma 3.6 we obtain for any
Since the adjoint operator leaves invariant, we may iterate the above argument to obtain
Since was arbitrary we have
(3.7) | ||||
We now complete the proof of existence of the map . Suppose that a certain linear combination of vectors of the form vanishes. That is, suppose we have
for , quasiprimary vectors with conformal dimension , and Laurent polynomials. Then, by the above, the function
(3.8) |
extends holomorphically to a neighborhood of , and by (3.7) the derivatives of (3.8) at are given by
Since the Taylor series of (3.8) at is identically zero, the function vanishes identically. In particular, specializing to yields
We conclude that the desired map is well-defined, as required. ∎
We now address the problem of extending to an endomorphism of .
Lemma 3.8.
Let be a Möbius vertex algebra which is generated by its quasiprimary vectors. Then for any there exists a unique linear map such that
for all with conformal dimensions and all .
Proof.
Let , and recall that . Thus we may canonically write with (and the sum converging in the weak topology induced by ). We would like to define , but first must check convergence of the sum. It suffices to consider a vector with and . Recall from Lemma 3.2 that the continuous map multilinear map given by
takes values in with a finite-dimensional subspace of . Thus by the universal property of the projective tensor product [Trè67, Prop. 43.4] we have a continuous linear map
As is complete, by [Trè67, Thm. 5.2] (see also Appendix B for the completion of topological vector spaces) we may extend this map to a continuous linear map
where is the completed projective tensor product (see [Trè67, §43]). We have a natural isomorphism of topological vector spaces
by [Trè67, Thm. 56.1] (extended to the manifold via partition of unity). Thus we conclude that there exists a continuous linear map
characterized by
Now fix and , and consider
again with . Let be the continuous linear map such that
for all . For , let be the natural projection onto the closed span of monomials with (whose kernel is spanned by monomials with ). For , by construction we have and , where is defined in Lemma 3.7. Since (with convergence pointwise as operators on ) and both and are continuous, we have convergence of the sum
to .
As the action of on is well-defined by Lemma 3.7 and does not depend on the choice of , we have obtained both a well-defined map given by along with the required covariance relation. ∎
Since , the maps furnish a representation of on . For the operator-valued distribution is evidently covariant with respect to this representation.
Theorem 3.9.
Let be a Möbius vertex algebra, let be a set of quasiprimary vectors that generate as a vertex algebra, and let
Let and for let . Then is a (not-necessarily-unitary) Möbius-covariant Wightman CFT.
Proof.
We have a family of operator-valued distributions on . Note that since generates . By Lemma 3.4 we have , where we note that it suffices to check continuity of in the special case since is generated from by . Hence separates points, as separates points in , and so acts regularly. The subspace is invariant under by Lemma 3.8, and by the same lemma the fields are Möbius covariant, which verifies the first axiom of a Wightman CFT.
We now check the locality axiom. Let , let and let . By the vertex algebra locality axiom, the formal distribution vanishes for sufficiently large, and thus the corresponding distribution is supported on the diagonal (see [Kac98, Cor. 2.2] and [CKLW18, Prop. A.1]). Hence when and have disjoint support we have
That is,
for all and . By the joint continuity of such expressions in (which shows that is -weakly dense in ) and the cyclicity of , we see that for all , and thus the locality axiom holds.
The vacuum axiom holds by construction, and the spectrum condition holds by Lemma 2.11. ∎
3.2 From Wightman CFTs to vertex algebras
Let be a Wightman CFT with domain , with vacuum vector and representation . Let be the finite energy subspace
where , and let . Note that when we have by the spectrum condition of a Wightman CFT, and is -strongly dense in .
We will show that carries the structure of a Möbius vertex algebra generated by the point-like quasiprimary fields corresponding to . For with conformal dimension , the corresponding point-like field is a formal sum
The key steps are to establish the vertex algebra locality condition
for sufficiently large, as well as differentiating the representation to a representation of for which we have the infinitesimal Möbius covariance condition
From there, we will invoke general results that say that families of covariant local fields produce vertex algebras (see [Kac98, Thm. 4.5] for the case of vertex algebras, or more specifically [RTT22, Thm. A.1] for a slight variant for Möbius vertex algebras).
We begin by establishing Möbius covariance.
Lemma 3.10.
There is a unique representation such that for all with conformal dimension and all we have and
Proof.
Uniqueness of such a representation follows immediately from the cyclicity of the vacuum (W4). Let , and let be the associated one-parameter group. We have
where is the conformal dimension of . The derivative of is given (as in [RTT22, Eqn. (3.4)]) by
with the derivative taken in the space of smooth functions on .
Give the -strong topology. Since expressions are jointly continuous in the , we have
(3.9) |
In particular, for every the expression is differentiable at , and we define . We have by the Möbius invariance of the vacuum, and from (3.9) we obtain the desired commutation relation for . A direct calculation shows that is a Lie algebra representation. ∎
Recalling that for , one can apply Lemma 3.10 term-by-term to the modes of to deduce the infinitesimal covariance relation
(3.10) |
We now turn our attention to establishing the vertex algebra locality condition. Recall that denotes ; that is, the space of linear functionals on that are supported on finitely many . By abuse of notation we write for the subspace of consisting of linear functionals such that , and similarly for . By Lemma 2.11 separates points in , and so separates points in . The endomorphism of (3.10) and Lemma 3.10 gives an endomorphism of by the lemma. Moreover, the adjoint (transpose) operator leaves invariant, becauase if , then we have by the same lemma
which depends continuously on the , so . Hence leaves invariant, mapping into .
If is a linear functional on a vector space and is a formal series with coefficients in , then we write for the corresponding formal series with coefficients in .
Lemma 3.11.
Let with conformal dimensions and , respectively. Then for every the formal series
is a polynomial in and after expanding using the binomial theorem.
Proof.
We use standard vertex algebra arguments which go through in the present context. From the positivity of the energy and the - and -commutation relations (3.10), we can deduce (as in the proof of [RTT22, Thm. 3.11]) that has only non-negative powers of , and if is the constant term, then . The formal power series and both solve the initial value problem with . This initial value problem has a unique solution in , and we conclude as formal series.
Similarly, we consider the formal series in and given by . It satisfies the initial value problem with . Taking each coefficient of separately, it is straightforward to see that this initial value problem has a unique solution in . Let denote the series in obtained by expanding each term as a binomial series with positive powers of . This series satisfies the same initial value problem, and so we have
Putting the two calculations together, we obtain an identity of formal series
Hence
As maps into , it acts nilpotently on and the sum defining is finite.
Consider a term of this sum, which is of the form . It suffices to prove the lemma for and then take linear combinations, in which case there is at most one non-zero term in the sum defining this expression. That is, if we write then
Since this term is non-zero only when , we have that
for a constant , which is a polynomial in and . We conclude that
is a polynomial in and , and we are done. ∎
Lemma 3.12.
Let be a Möbius-covariant Wightman CFT, and let . Then and are local in the sense of vertex algebras.
Proof.
Let be the operator-valued distribution corresponding to the formal series after expanding out the binomial . More precisely, we first define on pairs of functions , where , by
and these coefficients lie in . However expanding we see that is a (finite) linear combination of distributions of the form
(3.11) |
which extends to a genuine distribution as claimed. Moreover, we see from this formula that when and have disjoint support, i.e. the support of is contained in the diagonal of .
Let , and note that since the distribution
is indeed continuous in and . Applying Lemma 3.11 twice, we see that this distribution, which corresponds to the formal series , is given by integration against a trigonometric polynomial. As noted above this distribution (and hence the corresponding polynomial) has support contained in the diagonal of , and thus must be identically zero. Since separates points in by Lemma 2.11 we conclude that for all . As is -weakly continuous in and , this implies that for all .
Recall that is a linear combination of distributions of the form (3.11). Hence if and are supported in an open, non-dense interval of the circle, then the Reeh-Schlieder property (Corollary A.3) implies that . Now choose three intervals that cover such that the union of any two is contained inside some interval, and let be a partition of unity subordinate to this cover. Then for arbitrary . In particular for all , and we conclude that the formal series is identically zero, as desired. ∎
We can now state and prove one of our main results, constructing a Möbius vertex algebra from a Wightman theory.
Theorem 3.13.
Let be a (not-necessarily-unitary) Möbius-covariant Wightman CFT on with domain , and let be given by
Then there is a unique structure of Möbius vertex algebra on such that for every with conformal dimension there is a quasiprimary such that . The set generates .
Proof.
We equip with the representation of from Lemma 3.10. To show that the point-like fields generate a Möbius vertex algebra, we invoke [RTT22, Thm. A.1] (see also [Kac98, Thm. 4.5]). To invoke his theorem, we need to verify that:
-
1.
-
2.
is -invariant
-
3.
For every , has a removable singularity at
-
4.
For every , there exists a such that
-
5.
For every , we have for sufficiently large
-
6.
.
The first point follows from the fact that is an eigenvector for with eigenvalue by the commutation relation of Lemma 3.10. The second point and fourth point also follow from Lemma 3.10 along with Equation (3.10). The fifth point holds by Lemma 3.12, and the sixth point is the definition of .
We now argue the third point, that has a removable singularity at . The argument is the same as in [RTT22, Thm. 3.11]. Let .333Note that is the -th mode of a single field and not the -th field. We use this notation only here and in the next paragraph.
We must show that for . When this identity holds by the spectrum condition which implies that for these . So we now consider . From the -commutation relation of we have
We repeatedly apply this identity, starting with , to obtain , as desired.
Thus by [RTT22, Thm. A.1] there exists a unique structure of Möbius vertex algebra on , with the same , such that for every with conformal dimension we have . The vector is quasiprimary, as
By the sixth point, the set in the statement of the theorem generates . This completes the proof of the existence statement.
For uniqueness, note that the set generates any vertex algebra satisfying the statement of the theorem. The modes of the corresponding fields are determined by the fields , and the modes of the remaining fields are then determined by the Borcherds product formula (2.5). The grading operator is determined by the requirement that the conformal dimension of matches the conformal dimension of . The operators are then determined by the commutation relations with the generating fields. We conclude that the Möbius vertex algebra constructed above is the unique such structure satisfying the requirements of the theorem. ∎
As a corollary of the proof of Theorem 3.13 we have that if is non-zero and has conformal dimension then
(3.12) |
and in particular this gives a proof that the conformal dimension of a Wightman field is uniquely determined.
We conclude this section with a canonical realization of the domain of a Wightman CFT.
Proposition 3.14.
Let be a Möbius-covariant Wightman CFT on with domain , and let be the corresponding Möbius vertex algebra from Theorem 3.13. Equip with the -strong topology and equip with the weak topology induced by . Then the identity map extends to a (necessarily unique) injective continuous linear map .
Proof.
First, we claim that for any there exists a unique sequence such that , converging in the -strong topology. We first consider existence. It suffices to establish existence for . Arguing as in the proof of Lemma 3.8, there exists a continuous map such that
for all , where is the completion of in the -strong topology (see Appendix B) and
Let be the projection onto the closed span of monomials with (whose kernel is spanned by monomials with ). When we have
where is the vector corresponding to . Since has finite dimension, and is therefore complete [NB11, Thm. 4.10.3], the composed map takes values in , and in particular in . Thus if and we set , then in (the natural extension of) the -strong topology on , because in and is continuous.
We now consider uniqueness of the sequence . Suppose that with and the sum converging -strongly. Then any extends to by continuity (see Appendix B) and we have
As separates points in by Lemma 2.11 we see , and since was arbitrary this establishes the uniqueness portion of the claim.
We now define by , where is the unique sequence such that with -strong convergence. This map is well-defined by the above claim and, by inspection, is injective and restricts to the identity on . It remains to check that is continuous from the -strong topology to the weak topology on induced by . By the universal property of the -strong topology, it suffices to check that depends continuously on the for any . By the calculation above we have
(3.13) |
We have seen that is a continuous map with values in the finite-dimensional space , and is evidently continuous. We conclude that (3.13) is continuous in the , and so is continuous as claimed. ∎
3.3 Equivalence of categories
We have constructions in Theorem 3.9 and Theorem 3.13 that produce Wightman CFTs from vertex algebras and vice versa. In this section we show that these constructions are inverse to each other, or more precisely we show that they induce an equivalence of categories. We now introduce the relevant categories.
A homomorphism of Möbius vertex algebras is a linear map that intertwines the representations of , maps the vacuum vector to the vacuum vector, and intertwines the modes:
Now suppose that have are equipped with choices of generating sets of quasiprimary vectors and , respectively. We say that is a morphism if is a homomorphism of Möbius vertex algebras and . We write for the category of Möbius vertex algebras equipped with a choice of generating set of quasiprimary vectors.
If and are Möbius-covariant Wightman CFTs on , then a morphism is a linear map and a function such that , intertwines and , and for all and . Note that is uniquely determined by . A straightforward calculation shows that a homomorphism is continuous when and are respectively given the -strong and -strong topologies, and similarly for the -weak and -weak topologies. We write for the category of Möbius-covariant Wightman CFTs on .
Lemma 3.15.
Let and be a pair of Möbius-covariant Wightman CFTs and let be a morphism . Let and be the Möbius vertex algebras constructed in Theorem 3.13, and let and be the respective sets of generating vectors. Then and is a morphism in .
Proof.
By definition is spanned by vectors of the form where and . Since is a morphism we have
so .
We next check that intertwines the representations of . Let and let be the corresponding one-parameter group. We saw in the proof of Lemma 3.10 that the representations of on and are given by differentiating , and so we have
where we used that the derivatives are taken in the - and -weak topologies, and is continuous with respect to these topologies.
Now fix with conformal dimension . Let be the conformal dimension of , and we begin by arguing provided . By (3.12) we have
As we have , and we must show that . From the previous step we know that , where as usual and similarly for . Thus for we have
If , we may apply the above relation repeatedly to to conclude that for all . But then we would have , a contradiction. We conclude that , which is to say that and have the same conformal dimension provided .
Next observe that , or equivalently . We therefore have
This means that intertwines the actions of modes of vectors corresponding to , and since such vectors generate we can conclude that intertwines the actions of modes for all . Moreover the identity implies that . ∎
Lemma 3.16.
Let and be Möbius vertex algebras with generating sets and , respectively. Let be a morphism in . Let and be the Möbius-covariant Wightman CFTs constructed in Theorem 3.9. Then there is a unique morphism such that .
Proof.
For , we write for the corresponding Wightman field in , and similarly for we write for the Wightman field in . For , we define . Since is a morphism of vertex algebras we have for all and all
Since morphisms of Wightman CFTs are continuous for the - and -weak topologies, we can see from the above formula that a morphism as in the statement of the lemma is necessarily unique.
Since intertwines the actions of and , the adjoint operator maps into . As and (by Lemma 3.4), and separates points in , it follows that the closure of the graph of in is again the graph of a densely defined linear map . If , we may approximate each by Laurent polynomials to obtain
Hence is defined on all of and for all and . It follows immediately that also intertwines the representations and , and we have shown that is a morphism . ∎
Lemmas 3.15 and 3.16 upgrade the constructions of Theorem 3.13 and 3.9 to a pair of functors and . In showing that these are an equivalence of categories, it will be helpful to note that if is a Wightman CFT with domain , then the vertex algebra is a subspace . Conversely, if and is the domain of the Wightman CFT , then .
Lemma 3.17.
We have the following.
-
i)
Let be a Möbius-covariant Wightman CFT, let with . Let with . Then there is a unique isomorphism such that .
-
ii)
Let be a Möbius vertex algebra, let be the corresponding Möbius-covariant Wightman CFT with . Let be the Möbius vertex algebra . Then as Möbius vertex algebras.
Proof.
We first consider (i). Uniqueness of such an isomorphism follows from the fact that an isomorphism is -strong continuous and is -strong dense. We now consider existence. By construction there is a canonical bijection which we denote by . We must verify that there exists a corresponding bijection . We have and . By construction we have , and Proposition 3.14 provides a map . We have when , and since both sides are continuous in the functions this extends to all . We conclude that maps into and furnishes the necessary bijection. Part (ii) is immediate from the construction. ∎
The isomorphisms from Lemma 3.17 are natural in and respectively, and thus we have proven the following.
Theorem 3.18.
Let be the category of (not-necessarily-unitary) Möbius-covariant Wightman CFTs and let be the category of Möbius vertex algebras equipped with a generating family of quasiprimary vectors. Let be the functor constructed on objects in Theorem 3.13 and on morphisms in Lemma 3.15. Let be the functor constructed on objects in Theorem 3.9 and on morphisms in Lemma 3.16. Then and , along with the isomorphisms of Lemma 3.17, are an equivalence of categories between and .
4 Invariant forms and unitary theories
In this section we show that the correspondence between Wightman CFTs on and Möbius vertex algebras constructed in Section 3 is compatible with invariant bilinear forms. The definition of an invariant bilinear form for a Möbius vertex algebra is standard (see [FHL93, §5.2] and [Li94]).
Definition 4.1.
An invariant bilinear form on a Möbius vertex algebra is a bilinear form such that
(4.1) |
and
(4.2) |
for all .
It can be convenient to introduce notation for the opposite vertex operator
(4.3) |
and in this notation the invariance condition becomes
The map extends linearly to a Lie algebra automorphism of which leaves invariant. Let be this restriction. In this notation, the compatibility condition (4.2) between the invariant bilinear form and the representation of on becomes
In order to formulate the correct notion of invariant bilinear form for a Wightman CFT, we must integrate to an automorphism of . It is straightforward to check that is given by
Indeed, at the level of matrices is given on (with ) by complex conjugation
and is given on (with ) by complex conjugation as well
In particular we have
(4.4) |
We thus have the following notion of invariant bilinear form for a Wightman CFT.
Definition 4.2.
Let be a Möbius-covariant Wightman CFT on . A jointly -strong continuous bilinear form on is called an invariant bilinear form if
(4.5) |
for all (with conformal dimension ), all , and all , and moreover
(4.6) |
for all and .
As in the context of vertex algebras, we can introduce the notion of opposite field
and the invariance condition (4.5) then becomes
Theorem 4.3 (Correspondence between invariant bilinear forms).
Let be a Möbius-covariant Wightman CFT on and let be the corresponding Möbius vertex algebra. Then
-
i)
Every invariant bilinear form for the Wightman CFT restricts to an invariant bilinear form for the vertex algebra .
-
ii)
Every invariant bilinear form for the vertex algebra extends uniquely to an invariant bilinear form for the Wightman CFT on .
If an invariant form on is nondegenerate, then so is the extension to . Conversely, if an invariant form on is nondegenerate then so is the restriction to .
Proof.
First suppose that is equipped with an invariant bilinear form . Let , let , and let
For we have
Differentiating and evaluating at (as in the proof of Lemma 3.10) yields
as required. Now let be the set of quasiprimary generators corresponding to . For we have
and in particular at the level of modes we have
Hence for we have
This extends to all by Lemma 4.4 below, and we have established (i).
Now conversely suppose that is equipped with an invariant bilinear form which we denote . Note that a -strongly continuous extension of such a form on to a bilinear form on is unique, and so we must only show existence. Recall from Proposition 3.14 that comes naturally embedded in . The bilinear form on naturally extends to a pairing of and . First, we claim that for , and we have
(4.7) |
Indeed these agree when since the form is invariant for , and this identity extends to all smooth functions by continuity.
With this in mind, we wish to define a bilinear form on by extending linearly the prescription
(4.8) |
but we must first check that this is well-defined. Suppose that for some collection of Wightman fields and smearing functions we have
Then for all and
and thus
This shows that the prescription (4.8) is well-defined in the first input. We may repeat the above argument (invoking (4.7)) to show that it is also well-defined in the second input, and we conclude that (4.8) extends to a well-defined bilinear form. As (4.8) is continuous in the functions and , the bilinear form on is jointly -strong continuous, as required.
Finally, we need to check that is compatible with the representation of . Let , let and recall that . From the proof of Lemma 3.6 we have for
with the derivative taken in the -strong topology on . Similarly
Hence by the joint continuity of the bilinear form we have
In the last equality we used the fact that for , which extends to vectors in by the -strong continuity of and . Hence the above expression is independent of , and as the exponential map is surjective we then have
for all and , as required.
Finally, we address nondegeneracy. Recall that embeds naturally in and that the bilinear form on is compatible with the pairing of and . If the bilinear form on is nondegenerate and (with ) is non-zero, then for some , and thus there exists such that
A similar argument shows that the right-kernel of the form is zero, and so the form on is nondegenerate.
Conversely, assume that the form on is nongenerate. By Möbius invariance, its restriction to is the direct sum . Let with . Then there exists with
Hence there must be some such that , and so the left-kernel of the form on is zero. A similar argument shows that right-kernel is zero as well. ∎
We used the following fact in the proof of Theorem 4.3.
Lemma 4.4.
Let be a Möbius vertex algebra equipped with a bilinear form , and let be a set of vectors that generate . Suppose that the invariance condition
holds for and , and also that
for all . Then the invariance condition holds for all (i.e. the form is an invariant bilinear form for ).
Proof.
There is a (generalized) -module structure on the restricted dual whose state-field correspondence is characterized by
for all and . This contragredient module structure was first studied in [FHL93, §5.2] and described further in our context with infinite-dimensional weight spaces in the paragraphs following [HLZ14, Lem. 2.22]. If is the map , then our hypothesis implies that for all , or at the level of modes for all and . This intertwining condition extends to all by the Borcherds product formula (for and for ), and we conclude that the bilinear form is invariant. ∎
It was shown in [FHL93, Prop. 5.3.6] that every nondegenerate invariant bilinear form on a vertex operator algebra is symmetric. Later it was observed in [Li94, Prop. 2.6] that the proof does not use the hypothesis of nondegeneracy, and further examination of the proof in [FHL93] shows that the proof also goes through for Möbius vertex algebras as defined in this article (that is, allowing for infinite-dimensional -weight spaces and only using Möbius symmetry rather than Virasoro). In light of Theorem 4.3, we have the same result for Wightman CFTs.
Corollary 4.5.
Every invariant bilinear form on a Wightman CFT is symmetric.
We now turn our attention to unitary theories, and more generally invariant sesquilinear forms (which we call involutive structures). In order to do this we will need to introduce antilinear homomorphisms of Möbius vertex algebras and Möbius-covariant Wightman CFTs. Let and be Möbius vertex algebras, with vacuum vectors and and representations and of , respectively. Then an antilinear map is called a homomorphism if and
for all , and .
On the Wightman side, if and are Möbius-covariant Wightman CFTs, an antilinear homomorphism is an antilinear map and a function such that and
for all , and (and we recall )444It may be surprising that the condition for vertex algebras corresponds to for Wightman CFTs. Note that due to the antilinearity of , the relation does not imply that intertwines the representations of . In fact we have for .. We have
where denotes the pointwise complex conjugate. Just as we demonstrated in Lemmas 3.15 and 3.16, one can show that antilinear homomorphisms of Möbius vertex algebras extend uniquely to antilinear homomorphisms of Wightman CFTs, and conversely antilinear homomorphisms of Wightman CFTs restrict to antilinear homomorphisms of Möbius vertex algebras.
Lemma 4.6.
Let and be two Möbius-covariant Wightman CFTs and let and be the corresponding Möbius vertex algebras with respective generating sets and .
-
i)
If is an antilinear homomorphism then and is an antilinear homomorphism of Möbius vertex algebras satisfying .
-
ii)
If is an antilinear homomorphism of Möbius vertex algebras such that , then there is a unique antilinear homomorphism such that .
We omit the proof of Lemma 4.6 which is essentially identical to Lemma 3.15 and 3.16 (once we observe as above that an antilinear vertex algebra homomorphism satisfies for ).
Recall that an antilinear map is said to preserve a sesquilinear form if for all vectors , and that a sesquilinear form is said to be (Hermitian) symmetric if .
Definition 4.7.
An involutive Möbius vertex algebra is a Möbius vertex algebra equipped with a sesquilinear form and an antilinear automorphism which is involutive and preserves the sesquilinear form, and such that is an invariant bilinear form. An involutive Möbius vertex algebra is called unitary if the sesquilinear form is an inner product that is normalized so that .
We use the convention that sesquilinear forms are linear in the first variable, and require that homomorphisms of Möbius vertex algebras commute with the operators . The condition that is an invariant bilinear form is equivalent to having
(4.9) |
for all and , where is a formal complex variable, i.e. .
We sometimes refer to the sesquilinear form from Definition 4.7 as an invariant sesquilinear form, omitting reference to the involution .
Remark 4.8.
The sesquilinear forms from Definition 4.7 are automatically Hermitian symmetric as a consequence of the fact that invariant bilinear forms are symmetric. If the sesquilinear form is nondegenerate then the requirement that be involutive is redundant and the automorphism is uniquely determined by the sesquilinear form (the proof is exactly as in [CKLW18, Prop. 5.1] for inner products).
We now turn our attention to invariant sesquilinear forms on Wightman CFTs.
Definition 4.9.
An involutive Möbius-covariant Wightman CFT on is a Wightman CFT along with a jointly -strong continuous sesquilinear form on and an involutive automorphism of such that preserves the sesquilinear form and such that is an invariant bilinear form. An involutive Möbius-covariant Wightman CFT is called unitary if the sesquilinear form is an inner product which is normalized so that .
As with vertex algebras, we sometimes refer to the sesquilinear form of Definition 4.9 as an invariant sesquilinear form, omitting reference to the involution.
If we write , then the condition that is an invariant bilinear form is equivalent to
(4.10) |
for all , , and 555Note that this is a slight departure from [RTT22], where we required that be invariant under the involution rather than . In the present setting we find the updated definition to be more natural, as the involution of fields typically does not correspond to an antilinear automorphism of the Wightman CFT.. Here, as before, denotes the pointwise complex conjugate of the function .
As with involutive vertex algebras (Remark 4.8), the sesquilinear form of an involutive Wightman CFT is automatically Hermitian symmetric.
Theorem 4.10 (Equivalence of involutive and unitary structures).
Let be a Möbius-covariant Wightman CFT, and let be the corresponding Möbius vertex algebra equipped with a set of quasiprimary generators. Then we have the following.
-
i)
If is equipped with a sesquilinear form and involution making it into an involutive Wightman CFT, then the sesquilinear form and involution restrict to an involutive structure on the vertex algebra . The set of quasiprimary generators is invariant under .
-
ii)
If is equipped with a sesquilinear form and involution making it into an involutive vertex algebra and is invariant under , then there is a unique involution of and unique extensions of the sesquilinear form and to making into an involutive Wightman CFT.
If the sesquilinear form is nondegenerate on then it remains nondegenerate on , and similarly if the form is nondegenerate on so is the extension to . Moreover unitary structures on correspond to unitary structures on , and vice versa.
Proof.
First consider an involutive structure on . Then restricts to an antilinear involutive Möbius vertex algebra automorphism by Lemma 4.6. By Theorem 4.3, the invariant bilinear form restricts to an invariant bilinear form on , and it follows that and yield an involutive structure on . If corresponds to the state , then
Hence the Wightman field corresponds to and is -invariant (we have used here the observation that preserves the conformal dimension of fields).
For the other direction, suppose that we have an involutive structure on corresponding to an involution and sesquilinear form . Then the invariant bilinear form extends uniquely to an invariant bilinear form on . By Lemma 4.6 we may uniquely extend to an antilinear automorphism of . This extension is -strong continuous, and thus the sesquilinear form is -strong continuous as well. This sesquilinear form, along with , yield an involutive structure on as required.
The proof of equivalence of nondegeneracy is straightforward (as in the proof of Theorem 4.3), and the equivalence of unitarity is immediate. ∎
For unitary Wightman CFTs domain can be equipped with the norm topology coming from the inner product. This leads to a number of analytic questions, which are discussed further in [RTT22].
Remark 4.11.
Suppose that is a Wightman CFT such that is equipped with an inner product and is equipped with an involution . The definition of a (unitary) Wightman CFT given in [RTT22] required only that the compatibility conditions (4.10) hold, with no mention of the PCT operator . However, under these assumptions one may show that there exists a unique making the associated vertex algebra into a unitary vertex algebra, arguing as in [RTT22, Thm. 3.11] based on [CKLW18, Thm. 5.16]666The hypothesis that required in [CKLW18, Thm. 5.16] is not needed, as shown in [CGH23, §3.4], and the condition is also not needed.. One may then extend to an antilinear involution of by Lemma 4.6, making into a unitary Wightman CFT as defined in this article. We note that for a general sesquilinear form on , an involution satisfying (4.10) does not necessarily correspond to an involutive structure. Indeed, in the extreme example where the sesquilinear form is identically zero, the compatibility conditions (4.10) impose no constraint on the involution , but not every set-theoretic involution of corresponds to an involutive structure. It is possible that the conditions (4.10) are sufficient to reconstruct the PCT operator when the sesquilinear form is nondegenerate, but we do not address that question here.
Appendix A The Reeh-Schlieder theorem for non-unitary
Wightman conformal field theories
In this section we work with rotation-covariant Wightman CFTs on , which differ from Möbius-covariant Wightman CFTs (Definition 2.9) only in that the symmetry is only a representation of of the rotation subgroup , and accordingly the covariance condition (W1) is weakened to only require covariance for rotations. We write either or for rotation by .
For an interval (i.e. is a connected open non-empty proper subset), we let be the algebra generated by smeared fields with . The goal of this section is to establish the Reeh-Schlieder theorem for the theory , which says that the vacuum vector is cyclic and separating for the algebras . Recall that is cyclic for an algebra (with respect to a certain topology on ) if is dense in , and separating for if the only such that is . The analogous statement for unitary Wightman quantum field theories on higher-dimensional spacetimes is well-known (see [SW64, §4.2] and [RS61]). We give here a proof of the Reeh-Schlieder theorem in our current (not necessarily unitary) context.
Let be the open unit disk in , and its closure. We denote by the space of continuous -valued functions on that are holomorphic on the interior . By the maximum principle embeds as a closed subspace of , and we give the norm inherited from .
Lemma A.1.
Let be a rotation-covariant Wightman CFT. Fix , and let . Let and let . Then for each , the map
(A.1) |
lies in .
Proof.
When the functions are all Laurent polynomials, the expression (A.1) is a polynomial in the and the conclusion follows. We now consider the general case.
By rotation covariance we have
where , and is the conformal dimension of . Given arbitrary smooth , choose sequences of Laurent polynomials such that in . As in Section 3.1, let be the Sobolev space corresponding to a number , and recall that the topology on is generated by the Sobolev norms . Since acts as a unitary on , we have convergence in each
that is uniform in .
Lemma A.2.
Let be a rotation-covariant Wightman CFT on with domain , and let be an interval. Let , and suppose for all . Then .
Proof.
Fix , so that
(A.3) |
whenever for . Fix supported in , and consider the function given by
We have by Lemma A.1. Moreover, by rotation covariance vanishes on a small interval of about (note that is closed and the interval is open, so that contains a neighborhood of ). Thus by the Schwarz reflection principle we have identically, and restricting to we have
for all . Hence (A.3) holds whenever are supported in , and is supported in any interval of length . Using a partition of unity, it follows that (A.3) holds for arbitrary .
We now repeat the above argument. As before, we may show that the function
vanishes identically on , and from there deduce that (A.3) holds whenever are supported in , and are arbitrary. Repeatedly applying this argument, we see that (A.3) holds for all , which means by the vacuum axiom of a Wightman CFT. ∎
Corollary A.3 (Reeh-Schlieder theorem).
Let be a rotation-covariant Wightman CFT on with domain . For an interval we let be the subalgebra generated by with and . Then
-
i)
is cyclic for with respect to the -strong topology on , i.e. is -strongly dense in .
-
ii)
is separating for , i.e. if and then .
Proof.
For part (i), recall from Remark 2.6 that is precisely the dual space of equipped with the -strong topology. By Lemma A.2 the closed subspace is annihilated only by the zero functional, and so by the Hahn-Banach theorem (for locally convex topological vector spaces) we must have .
For part (ii), observe that by the locality axiom of a Wightman theory the operator vanishes on , where is the interval complementary to . By Lemma 2.8 the operator is -strongly continuous, and hence by part (i) we have . ∎
Appendix B Topological vector spaces
In this section we supplement the discussion of the topology on the domain of a Wightman field theory by giving additional definitions, details, and references regarding topological vector spaces and locally convex spaces. We refer readers to the textbooks [NB11, Trè67] for further reading. All vector spaces in this section are assumed to be over the field of complex numbers.
A topological vector space is a vector space equipped with a vector topology, which is a topology such that the addition map and the scalar multiplication map are continuous. Vector topologies are not necessarily Hausdorff by definition, although we will primarily be interested in Hausdorff topological vector spaces.
A seminorm on a vector space is a map such that and for all and . Given a set of seminorms on , the corresponding seminorm topology is the coarsest topology on making all of the seminorms continuous. Seminorm topologies are always vector topologies, but not every vector topology is a seminorm topology. A locally convex space is a topological vector space whose topology is a seminorm topology corresponding to some set of seminorms. Equivalently, a locally convex space is a topological vector space such that there exists a neighborhood basis of the origin consisting of convex sets [NB11, Thm. 5.5.2]. Every Hausdorff topological vector space has a unique completion [Trè67, §5], and the completion of a locally convex space is locally convex [NB11, Thm. 5.11.5]. We note that finite-dimensional Hausdorff topological vector spaces are complete [NB11, Thm. 4.10.3], as are products of complete topological vector spaces. Every continuous linear map of Hausdorff topological vector spaces extends continuously to a map [Trè67, Thm. 5.2].
Locally convex spaces play an important role in functional analysis because the Hahn-Banach theorem holds for them. In particular, the continuous linear functionals on a locally convex Hausdorff space separate points. Moreover, if is a closed subspace of a locally convex Hausdorff space and , then there exists a continuous linear functional such that and [NB11, Thm. 7.7.7]. In contrast, there exist topological vector spaces which do not admit nonzero continuous linear functionals, such as spaces with .
Most familiar examples of topological vector spaces, such as normed vector spaces, are locally convex. Another source of locally convex spaces is via weak topologies [NB11, §8.2]. Given a vector space and a set of linear functionals on , the weak topology (or initial topology) on corresponding to is the coarsest topology making all of the functionals continuous. This is a locally convex vector topology, being the seminorm topology corresponding to the seminorms for . A sequence (or net) converges to if and only if for every . A map is continuous with respect to the weak topology if and only if is continuous for every .
Dually, we have the notion of the colimit (or final or strong) topology. Consider a vector space , and a family of linear maps from topological vector spaces such that the images span . The colimit topology on corresponding to the maps is the finest topology on such that every is continuous, and it is a vector topology [NB11, §4.11]. If is a topological vector space, then a linear map is continuous if and only if is continuous for all .
If each space is locally convex then we may define a subtly different locally convex colimit topology on , which is the finest locally convex topology such that each is continuous [NB11, §12.2]. If is a locally convex space then a linear map is continuous for the locally convex colimit topology if and only if is continuous for all [NB11, Thm. 12.2.2].
We now discuss tensor products of locally convex spaces. If ,, and are vector spaces then bilinear maps correspond to linear maps , where is the algebraic tensor product. If ,, and are locally convex spaces, then there is a unique locally convex topology on , called the -topology (or projective topology), such that jointly continuous bilinear maps correspond to continuous linear maps [Trè67, Prop. 43.4]. We write for the algebraic tensor product equipped with the topology.
We now conclude by revisiting the -strong topology. Suppose that is a set of operator-valued distributions on with domain a vector space . For every and we have a multilinear map given by . These correspond to linear maps
We include the case , in which case assigns . The -strong topology on is then defined to be the locally convex colimit of the maps . Unpacking the definitions, if is a locally convex space then a map is -strong continuous if and only if is jointly continuous in the for all and .
References
- [AGT23] M. S. Adamo, L. Giorgetti, and Y. Tanimoto. Wightman fields for two-dimensional conformal field theories with pointed representation category. Comm. Math. Phys., 404(3):1231–1273, 2023.
- [AMT24] M. S. Adamo, Y. Moriwaki, and Y. Tanimoto. Osterwalder–Schrader axioms for unitary full vertex operator algebras. arXiv:2407.18222 [math-ph], 2024.
- [BK08] G. Buhl and G. Karaali. Spanning sets for Möbius vertex algebras satisfying arbitrary difference conditions. J. Algebra, 320(8):3345–3364, 2008.
- [CGH23] S. Carpi, T. Gaudio, and R. Hillier. From vertex operator superalgebras to graded-local conformal nets and back. arXiv:2304.14263 [math.OA], 2023.
- [CKLW18] S. Carpi, Y. Kawahigashi, R. Longo, and M. Weiner. From vertex operator algebras to conformal nets and back. Mem. Amer. Math. Soc., 254(1213), 2018.
- [CWX] S. Carpi, M. Weiner, and F. Xu. From vertex operator algebra modules to representations of conformal nets. In preparation.
- [EG17] D. E. Evans and T. Gannon. Non-unitary fusion categories and their doubles via endomorphisms. Adv. Math., 310:1–43, 2017.
- [FHL93] I. B. Frenkel, Y.-Z. Huang, and J. Lepowsky. On axiomatic approaches to vertex operator algebras and modules. Mem. Amer. Math. Soc., 104(494):viii+64, 1993.
- [Gui20] B. Gui. Unbounded field operators in categorical extensions of conformal nets. arXiv:2001.03095 [math.QA], 2020.
- [Gui21] B. Gui. Categorical extensions of conformal nets. Comm. Math. Phys., 383(2):763–839, 2021.
- [Gui24a] B. Gui. Convergence of sewing conformal blocks. Commun. Contemp. Math., 26(3):Paper No. 2350007, 65, 2024.
- [Gui24b] B. Gui. Sewing and propagation of conformal blocks. New York J. Math., 30:187–230, 2024.
- [GZ23] B. Gui and H. Zhang. Analytic Conformal Blocks of -cofinite Vertex Operator Algebras I: Propagation and Dual Fusion Products. arXiv:2305.10180 [math.QA], 2023.
- [HLZ14] Y.-Z. Huang, J. Lepowsky, and L. Zhang. Logarithmic tensor category theory for generalized modules for a conformal vertex algebra, I: introduction and strongly graded algebras and their generalized modules. In Conformal field theories and tensor categories, Math. Lect. Peking Univ., pages 169–248. Springer, Heidelberg, 2014.
- [Hua99] Y.-Z. Huang. A functional-analytic theory of vertex (operator) algebras. I. Comm. Math. Phys., 204(1):61–84, 1999.
- [Hua03] Y.-Z. Huang. A functional-analytic theory of vertex (operator) algebras. II. Comm. Math. Phys., 242(3):425–444, 2003.
- [Hua20] Y.-Z. Huang. Generators, spanning sets and existence of twisted modules for a grading-restricted vertex (super)algebra. Selecta Math. (N.S.), 26(4):Paper No. 62, 42, 2020.
- [Kac98] V. Kac. Vertex algebras for beginners, volume 10 of University Lecture Series. American Mathematical Society, Providence, RI, second edition, 1998.
- [Li94] H. S. Li. Symmetric invariant bilinear forms on vertex operator algebras. J. Pure Appl. Algebra, 96(3):279–297, 1994.
- [Mor23] Y. Moriwaki. Two-dimensional conformal field theory, full vertex algebra and current-current deformation. Adv. Math., 427:Paper No. 109125, 74, 2023.
- [NB11] L. Narici and E. Beckenstein. Topological vector spaces, volume 296 of Pure and Applied Mathematics (Boca Raton). CRC Press, Boca Raton, FL, second edition, 2011.
- [RS61] H. Reeh and S. Schlieder. Bemerkungen zur Unitäräquivalenz von Lorentzinvarianten Felden. Nuovo Cimento (10), 22:1051–1068, 1961.
- [RTT22] C. Raymond, Y. Tanimoto, and J. E. Tener. Unitary vertex algebras and Wightman conformal field theories. Comm. Math. Phys., 395(1):299–330, 2022.
- [Seg04] G. Segal. The definition of conformal field theory. In Topology, geometry and quantum field theory, volume 308 of London Math. Soc. Lecture Note Ser., pages 421–577. Cambridge Univ. Press, Cambridge, 2004.
- [Str93] F. Strocchi. Selected topics on the general properties of quantum field theory, volume 51 of World Scientific Lecture Notes in Physics. World Scientific Publishing Co., Inc., River Edge, NJ, 1993.
- [SW64] R. F. Streater and A. S. Wightman. PCT, spin and statistics, and all that. W. A. Benjamin, Inc., New York-Amsterdam, 1964.
- [Ten19a] J. E. Tener. Geometric realization of algebraic conformal field theories. Adv. Math., 349:488–563, 2019.
- [Ten19b] J. E. Tener. Representation theory in chiral conformal field theory: from fields to observables. Selecta Math. (N.S.), 25(5):Paper No. 76, 82, 2019.
- [Ten24] J. E. Tener. Fusion and positivity in chiral conformal field theory. Geom. Funct. Anal., 34(4):1226–1296, 2024.
- [Trè67] F. Trèves. Topological vector spaces, distributions and kernels. Academic Press, New York-London, 1967.